Вы находитесь на странице: 1из 7

Modeling the Effects of Backfilling and Soil Compaction

beside Shallow Buried Pipes


Downloaded from ascelibrary.org by L & T Institute of Project Management on 12/25/14. Copyright ASCE. For personal use only; all rights reserved.

Tamer M. Elshimi, Ph.D., P.Eng 1; and Ian D. Moore 2

Abstract: Compaction of the soil placed beside culverts (the side-fill) can have a significant effect on the behavior of flexible and rigid
structures. This is particularly true for shallow buried structures when the stresses resulting from compaction represent a greater proportion
of the total stresses present. Different techniques have been reported in the literature to model soil compaction during finite-element
analyses. A new semiempirical technique is proposed, which takes into consideration the increase in lateral stress and soil kneading during
compaction. A simple procedure is discussed to incorporate the compaction of granular material in finite-element analysis. The new technique is used to model five different pipe products composed of different materials and dimensions, and the results are compared to
measured values reported in the literature. A new factor is proposed to account for soil kneading during compaction to provide an upper
limit for the pipe deformations and stresses that result during installation. The technique can be used to estimate peaking in flexible culverts
and the additional crown moments and thrust that result in rigid culverts. DOI: 10.1061/(ASCE)PS.1949-1204.0000136. 2013 American
Society of Civil Engineers.
Author keywords: Culvert; Backfill; Compaction; Modeling; Finite element analysis; Deflection; Thrust; Moment.

Introduction
Modeling the behavior of buried structures during construction requires the treatment of soil compaction. In particular, the effect of
compacting the soil placed beside the structure (the side-fill) can be
significant for shallow buried pipes and culverts (at deeper cover,
the response to the overburden soil dominates the pipe response).
Because the effect of compaction depends on the soilpipe interaction and characteristics such as pipe stiffness, pipe shape, and
backfill properties, there is no set range of burial depths in which
compaction is important; there is value in having computational
techniques to quantify this issue. Different techniques have been
reported in the literature for modeling this soil compaction. Duncan
and Seed (1986) developed a complex semiempirical procedure
to incorporate the stress path associated with soil compaction
within their nonlinear elastic soil model. This procedure, however,
has proven difficult for subsequent researchers and analysts to
apply and cannot be used in conjunction with elastic-plastic soil
models.
Work has recently been performed to develop explicit models
for soil compaction by using vibratory rollers (Kelm and Grabe
2004; Hgel et al. 2008), and for static compaction under individual
wheels [e.g., the work on tiresoil interaction by Hambleton and
Drescher (2009) and Xia (2011)]. However, these techniques seek
to explicitly model the geometrical path and impact of the wheel or
roller and will not be practical for modeling soil compaction beside
buried pipes, because:
1
Geotechnical Engineer, Thurber Engineering Ltd., 200, 9636-51
Avenue NW, Edmonton, AB, Canada T6E 6A5.
2
Professor and Canada Research Chair, GeoEngineering Centre at
QueensRMC, Queens Univ., Kingston, Ontario, Canada K7L 3N6
(corresponding author). E-mail: moore@civil.queensu.ca
Note. This manuscript was submitted on October 9, 2010; approved on
January 16, 2013; published online on January 18, 2013. Discussion period
open until January 16, 2014; separate discussions must be submitted for
individual papers. This paper is part of the Journal of Pipeline Systems
Engineering and Practice, ASCE, ISSN 1949-1190/04013004(7)/
$25.00.

ASCE

There are many layers to be placed and compacted beside the


pipe, and the analysis will be extremely time consuming; and
The pattern of movement of the equipment is not known in
advance, and will be difficult to record and simulate even for
post-construction analysis.
McGrath et al. (1999) summarized three other practical techniques that can be used to model compaction effects. One must
consider the modeling of a newly placed layer beside the culvert
[Fig. 1(a)]. The first technique is to apply a temporary surcharge on
the surface of that layer of backfill [a technique first proposed by
Katona (1978); Fig. 1(b)]. The second is to squeeze the most recently placed layer of backfill between two sets of temporary vertical pressures [Fig. 1(c)]. The third technique is to apply horizontal
nodal forces directly to the pipe [Fig. 1(d)]. McGrath et al. (1999)
concluded that:
The first two techniques have the advantage of creating temporary pipe distortion and deformations, but if used in conjunction
with elastic models, those deformations are lost when the temporary loads are removed;
The first two techniques are unsuccessful in creating deformations like those measured in the field for flexible pipes and were
unsuccessful in creating any significant peaking (when the
crown of a flexible pipe moves upwards during placement of
side-fill);
Use of the third technique requires the assessment of empirical
values for the lateral pressures that produce the correct peaking
of the pipe at the end of side-fill placement;
The lateral pressures back-calculated for metal and polymer
pipes work successfully for both pipe types, provided the diameter is the same;
Unfortunately, the lateral pressures depend on pipe diameter, so
the empirical exercise of calculating those pressures must be undertaken for each different diameter, soil type, and compaction
type under consideration; and
This third technique, therefore, requires a separate algorithm
or chart to provide guidance on the magnitude of the horizontal
forces that need to be applied to the culvert; McGrath et al.
(1999) determined values for three levels of compaction

04013004-1

J. Pipeline Syst. Eng. Pract. 2013.4.

J. Pipeline Syst. Eng. Pract.

Native Soil
Culvert
New placed layer

(a)

Downloaded from ascelibrary.org by L & T Institute of Project Management on 12/25/14. Copyright ASCE. For personal use only; all rights reserved.

Culvert
New placed layer

(b)

Culvert
New placed layer

(c)

Culvert

New placed layer

(d)

Fig. 1. Illustration of compaction models (described by McGrath et al.


1999): (a) new layer beside pipe; (b) surface stress on new layer;
(c) loads that squeeze the new layer; (d) application of horizontal loads
to flexible pipe

(rammer, vibratory plate, or none), just two soil materials


(stone and silty sand), and just two pipe diameters (0.9
and 1.5 m).
Other limitations of these models result because the methods are
designed for flexible pipe and do not work for concrete pipe structures, deep corrugated culverts (in which high flexural stiffness
means that they no longer respond like typical flexible culverts),
and other structures requiring calculations, like stormwater detention systems.
Taleb and Moore (1999) proposed a different technique to
model soil compaction for use in conjunction with elasticplastic
soil models. The method is simple and provides information on the
likely effects of soil compaction, both for flexible and rigid pipes.
The concept is to impose the high horizontal earth pressure that
remains in the soil immediately after compaction. This horizontal
earth pressure is set equal to the passive earth pressure (the maximum value that can be sustained by purely frictional soil). The responses of flexible and rigid structures to this additional lateral
earth pressure will be very different. Fig. 2 [adapted from the study
by Taleb and Moore (1999)] illustrates these two cases. For flexible
structures, the structure will deform under this horizontal pressure
and much of the additional lateral earth pressure will be released.
For rigid structures, the structure will not deform and the horizontal
earth pressures will remain, acting to change thrusts and moments,
especially at the crown.
ASCE

A simpler process is preferable, which can provide an upper


bound for the pipe deformations and stresses that can result from
compaction, for use in design projects for specific structures, and
parametric studies to develop more general design equations.
Although the procedure proposed by Taleb and Moore (1999)
satisfies many of these objectives, it neglects the plastic strains that
can be imposed during compaction (it applies the upper limits of
postcompaction stress associated with a purely frictional material,
but neglects the fact that kneading of the soil can lead to lateral
extension, i.e., plastic strains in the side-fill). The kneading produces higher horizontal expansion and vertical compression of the soil
layer than the time spent in compaction (or number of passes of the
compaction equipment) increases, but cannot produce additional
horizontal stresses because those are already at the limits dictated
by shear strength for purely cohesive soil. Therefore, the Taleb and
Moore (1999) technique may provide upper bounds to stresses that
develop on rigid structures, but will not necessarily produce upper
bounds to the movements that can occur in flexible pipes. Indeed,
Taleb and Moore (1999) used their technique to model the behavior
of a long-span corrugated metal culvert monitored during construction, and the procedure did not provide an upper bound to the structural deformation.
The Taleb and Moore (1999) technique employs the soil properties (shear strength, density, and modulus) for the added layer that is
applied at the end of compaction. This means that the effect on soil
properties of any densification or loosening of the soil as a result of
kneading (i.e., any hardening or softening) is included in the subsequent analysis steps (as further layers are added) That is, the consequences of hardening or softening are considered, although this
analysis approach does not include explicit modeling of the processes themselves. Therefore, this new study is focused on further
development of the simple procedure proposed by Taleb and Moore
(1999) to provide an upper bound to the effects of compaction and
to investigate its performance through comparisons with buried
pipe experiments. Specifically, the effect of soil kneading during
compaction will be taken into consideration by applying an initial
horizontal stress, which is equal to the passive earth pressure for
purely frictional material used by Taleb and Moore, multiplied
by an empirical kneading coefficient, K n . By using a nonzero value
of cohesion for the soil (e.g., 4 kPa), a value of K n up to 2 can be
accommodated in the new layer of soil at the ground surface without violating the failure criterion for frictional-cohesive soil; the
Mohr Coulomb criterion expressed in
pterms
of major and minor
principal stresses is 1 3 N 2c N , or for major principal
p
horizontal stress, h v N 2c N (cohesion, c; friction angle, ; and N 1 sin =1 sin ). By adding a small
amount of cohesion, c, the strength is increased relative to the
purely frictional case so that
p
2c N
h
N
> Kn N
v
v
That is, no shear failure when

p
N
v
c > K n 1
2

Elasticplastic analysis by using the Mohr-Coulomb failure criterion, a cohesion value of 4 kPa, and a maximum value for K n of 2
used in this study were sufficient to produce typical values of pipe
peaking and other desired responses for all of the cases under consideration. The horizontal earth pressures being imposed are those
associated with Rankine-Bell earth pressure equation, which is generally used to calculate stresses on smooth vertical earth retaining
structures. This choice is made to simplify the stress calculation and
is considered appropriate for an analysis that requires an empirical

04013004-2

J. Pipeline Syst. Eng. Pract. 2013.4.

J. Pipeline Syst. Eng. Pract.

Passive pressures in
newly placed layer

Culvert
Soil element

Native Soil

For rigid culvert, the lateral


movement is almost zero so
horizontal earth pressures
induce force and moment

Passive pressures
remain
Soil element

(b)
For flexible culvert,
horizontal stress on
the culvert causes it
to deform

Passive pressures
are then released
Soil element

(c)
Fig. 2. Illustration of the compaction model (after Taleb and Moore 1999): (a) imposed stress; (b) effect on rigid pipe; (c) effect on flexible pipe

factor like K n . Therefore, the effect on those pressures of shear


stresses on the soilpipe interface and any inclination of the sides
of the pipe are neglected.
This paper summarizes analysis performed to investigate the
effect of adding that small amount of cohesion and increasing horizontal stresses by the factor K n, considering five different pipes
manufactured with different materials and dimensions. Details
are presented about how to use this method in ABAQUS. The deformations, thrust forces, and bending moments were calculated by
using finite-element analysis; the displacements are compared to
the measured values for two different pipes reported by McGrath
et al. (1999), and to values measured for a 1.5-m-diameter highdensity polyethylene (HDPE) pipe backfilled with granular backfill
compacted with a vibrating plate. This procedure should provide an
upper bound to the peaking deformations for flexible pipes during
compaction. However, the analysis does not attempt to characterize
long-term effects associated with changes in moisture level in the
soil (either wetting or drying) or whether those horizontal stresses
are retained over the service life of the structure.

36, the vertical effective stress at the bottom of the layer will
be 18 0.3 5.4 kPa. If the initial conditions are not assigned,
the layer will deform under its own weight and the lateral stress
will be calculated by using the Poissons ratio.
Equilibrium is needed when the layer has an initial vertical
effective stress of 5.4 kPa at the bottom with zero deformations
at the top.
Under normal circumstances (if lateral stresses are not adjusted),
the lateral stress is calculated by using the earth pressure coefficient
(K o ) set equal to 1 sin (the common approach based on Jakys
approximation).
The Taleb and Moore (1999) procedure involves setting lateral
stresses to the passive values, so horizontal stress at the base of the
layer is equal to 20.8 kPa. Fig. 3 shows the distributions of vertical
Stress (kPa)
0.00
0
0.05
0.1

Incorporating the Compaction Technique in the


Finite-Element Analysis
Initial conditions are used in many finite-element analyses to
establish equilibrium and to ensure that displacements are zero after
applying geostatic stresses (when establishing the initial earth pressures). The initial stresses at two different depths are assigned and
the lateral earth pressure at rest, K o , is specified and used to calculate the lateral earth pressure. It is important to ensure that the
layer does not deform under its own weight.
For example, for a single layer of soil having a thickness of
0.3 m, a unit weight of 18 kN=m3 , and a friction angle () of
ASCE

Depth (m)

Downloaded from ascelibrary.org by L & T Institute of Project Management on 12/25/14. Copyright ASCE. For personal use only; all rights reserved.

(a)

-5.00

-10.00

-15.00

-20.00

-25.00

Vertical stress

(With compaction)
(Without compaction)

0.15
0.2
0.25
0.3
0.35

Fig. 3. Distribution of vertical and horizontal stresses in the soil layer


with and without the modeling of compaction effects

04013004-3

J. Pipeline Syst. Eng. Pract. 2013.4.

J. Pipeline Syst. Eng. Pract.

Downloaded from ascelibrary.org by L & T Institute of Project Management on 12/25/14. Copyright ASCE. For personal use only; all rights reserved.

and the horizontal stress versus depth, with and without modeling
of compaction effects. However, it was clear from the results of
Taleb and Moore (1999) that this technique did not provide an
upper limit for the peaking deformations resulting from soil compaction of the test structure they were studying. This is because the
technique neglected the effect of soil kneading during compaction
(in essence, the passive pressures can be applied and released by
culvert movements a number of times as the compaction process
occurs).
Table 1. Dimensions and Material Properties for Different Pipes
Dimensions
and properties RC Pipe 1 RC Pipe 2 CS Pipe 1 CS Pipe 2
2,000
2,000
Dinner (mm)
t 0 (mm)
190.5
100
190.5
100
A (mm2 =mm)
I P (mm4 =mm) 576,107.7 83,333.33
30,000
30,000
EP (MPa)

0.3
0.3
Sf
1.14
7.86

2,000
60.02
3.52
1,057.25
200,000
0.28
93.02

2,000
32.14
1.55
133.3
200,000
0.28
753.45

HDPE
pipe
2,000
63.26
9.44
3,111.1
450
0.46
1,0534.31

The effect of soil kneading can be approximated by adding the


empirical kneading factor (K n ). This increases the lateral stress to
(K n N v ), where v is the vertical stress. To determine the
value of K n , different values of K n have been used in the analysis
of five pipes with different material properties, dimensions, and
relative flexure stiffness (Sf ), defined by McGrath et al. (2002):
Sf M S R3 =EP I P

where M S = constrained soil modulus, a function of Youngs modulus, Es , and Poissons ratio, s , of the soil: M s Es 1 s =
1 s 1 2 s ; R = pipe radius; EP = Youngs modulus of
the pipe material; I P = second moment of area of the pipe/unit
length, which depends on the wall geometry (whether plain, corrugated, or of some more complex geometry).
The calculated deformations are compared to the measured
values reported by McGrath et al. (1999) in the next section.

Finite-Element Analysis of Five Different Pipes


Material Properties and Finite-Element Mesh

Table 2. Soil Parameters Used for SW95


Mean
confining Modulus of
Angle of Dilation
Unit
stress
elasticity Poissons friction: angle: Cohesion: weight
(kPa)
(MPa)
ratio:
()
()
c (kPa) (kN=m3 )
7
35
70
140

13.8
17.9
20.7
23.8

0.4
0.29
0.24
0.23

48
48
48
48

18
18
18
18

4
4
4
4

21
21
21
21

Note: Data from McGrath et al. (1999).

Table 3. Soil Parameters Used for SW85


Mean
confining Modulus of
Angle of Dilation
Unit
stress
elasticity Poissons friction: angle: Cohesion: weight
(kPa)
(MPa)
ratio:
()
()
c (kPa) (kN=m3 )
7
35
70
140

3.2
3.6
3.9
4.5

0.26
0.21
0.19
0.19

38
38
38
38

Note: Data from McGrath et al. (1999).

8
8
8
8

4
4
4
4

19
19
19
19

The behavior of five different pipe products during backfilling,


each having different relative flexural stiffness, Sf , was studied
by using two-dimensional (plane strain) finite-element analysis.
Table 1 shows the material properties and dimensions for these
pipes. Two different concrete pipes, two different corrugated steel
(CS) pipes (CSPI 2002), and one HDPE pipe were modeled. The
pipe dimensions and diameters were chosen to cover both rigid and
flexible pipes with different relative flexure stiffness values (Sf ), as
defined in Eq. (1).
Pipe installation was simulated in well graded sand compacted
to 95% of the maximum Proctor density (denoted SW95) for compacted soil, and well graded sand was compacted to 85% for loose
soil without compaction (SW85). Soil properties for SW95 and
SW85 are shown in Tables 2 and 3, respectively.
The soil was added in steps, each featuring a 200-mm layer
added to the model. Six node-modified quadratic plane strain triangle elements were used to model both the soil and the pipe. Fig. 4
shows the soil layers and the finite-element mesh for a reinforced
concrete pipe with diameter and thickness of 2 m and 190 mm,
respectively. The boundary conditions were applied at the axis
of symmetry and at both the bottom and side of the mesh. All
soilpipe interaction was modeled as bonded, and because the soil
modeling featured elasticplastic material response, the interface

Crown

Springline
Invert

Fig. 4. Soil layers modeled in the analysis and the finite-element mesh
ASCE

04013004-4

J. Pipeline Syst. Eng. Pract. 2013.4.

J. Pipeline Syst. Eng. Pract.

Results
Deformations
The technique described in the previous section was used to model
the compaction of well graded sand placed beside five different
pipe products. Fig. 5 shows the pipe deformations (changes in vertical and horizontal diameter, DV and DH , divided by the pipe
diameter, D) plotted against the Sf of each pipe for different values
of K n . The measured deformations reported by McGrath et al.
(1999) for two different pipes are shown on the same figure for
comparison with the calculated deformations. It is obvious that
the vertical and the horizontal deformations of the pipe during
the backfilling are dependent on Sf . Flexible pipes with Sf values
higher than 200 (Moore 2001), like HDPE pipes, deform during
backfilling much more than the rigid pipes with Sf values equal
to or less than 1, like reinforced concrete pipes (Moore 2001). Also,
the rate of increase in pipe deformations with the increase of Sf
starts to decrease at a certain point. As Sf value increases, the system eventually approaches the empty cavity case.
Fig. 5 shows that the magnitudes of vertical and horizontal
deformations of the pipes increase with an increase in K n . The increase in K n magnifies the lateral stress that is imposed in the soil
beside the pipe; therefore, it increases the vertical and horizontal
deformations (changes in pipe diameter). The deformations measured by McGrath et al. (1999) are effectively estimated when a
kneading factor of 2 is used, so it is proposed that a reasonable

upper bound to the pipe deformations can be obtained. This empirical value of K n matches the pipe response when a rammer is
used for the soil compaction. When a vibratory plate is used,
the measured deformations reported by McGrath et al. (1999) were
much smaller.
Fig. 6 shows that the compaction model captures the significant
effects on pipe deformations during side-filling. Compaction with a
rammer has a substantial effect on deformations in flexible pipes,
whereas compaction using a vibratory plate has a much lower effect. The vertical and the horizontal deformations of different pipes
calculated simulating compaction with K n of 2 are almost 10 times
the deformations calculated without modeling compaction (the
initial horizontal stress is calculated using the at rest lateral earth
pressure coefficient, K o ). This reflects how much additional deformation develops in the pipes and reflects the importance of
modeling the side-fill compaction, especially for shallow buried
structures. Figs. 79 show the deformed shapes of three different
pipes after backfilling up to the crown for different values of K n . It
is obvious that both the horizontal and peaking deformations are
higher when applying higher K n.
Thrust Forces and Moments
The bending moments and thrust forces at the springline and the
crown of the five different pipes were also calculated, and these are
plotted in Figs. 1013. First, it is clear that the effect of compaction
is not significant at the springline, although it changes the thrust
and moment at the crown. This is because the horizontal stresses
being imposed influence the horizontal force equilibrium; therefore, they thrust at the crown, rather than vertical force equilibrium
at the springline (Figs. 10 and 11). This is why compaction has a
significant effect on the bending moment at the crown for rigid
pipes (those having low Sf values). This results because the nonuniform pressure distribution around the pipe is increased because
of the horizontal stresses imposed by the compaction model. These
bending moments are approximately one-third of those that develop in rigid pipes as a result of one diameter of soil over the
crown, so shallow cover for rigid pipes may be defined as burial
depth of less than one diameter (so moments resulting from compaction are more than one-third of the total). For flexible pipes
(those having high Sf values), the pipe will deform under the imposed horizontal stresses, the pressure becomes more uniform, and
Sf

Sf
1

10

100

1000

10000

100

1000

10000

Dv/D

Dv/D (%)

10

Kn = 2

1
0

0
-1

-1

Dh/D

Dh/D (%)

Downloaded from ascelibrary.org by L & T Institute of Project Management on 12/25/14. Copyright ASCE. For personal use only; all rights reserved.

shear strength was limited to the shear strength of the soil.


Although this approach can be used when studying culvert response with elasticplastic soil models (Petersen et al. 2010), it may
provide conservative estimates of maximum hoop thrust and
unconservative estimates of moment and deformation (Moore
2001). For example, the Hoeg (1968) solution for pipes in purely
elastic ground indicates that by changing the interface on a buried
circular pipe from perfectly smooth (zero friction angle) to fully
bonded, changes in pipe diameter decrease by less than 14%,
and moment in rigid pipe also reduces by less than 14%. If a pipe
with an interface having a typical friction angle of 18 to 22 is modeled as fully bonded, these responses change by less than 5%. For
pipes bonded to elasticplastic soil, the changes are even smaller.

-2

-2
-3

Calculated, with compaction


Calculated, without compaction
Measured, Rammer
Measured, Vibratory plate
Measured, without compaction

-3

Calculated, Kn=2
Calculated, Kn=1.5
Calculated, Kn=1
Measured (McGrath 1999)

Fig. 5. Horizontal and vertical diameter change after side-filling (as a


function of Sf and K n )
ASCE

Fig. 6. Horizontal and vertical diameter change after side-filling as a


function of Sf ; comparison with measured values

04013004-5

J. Pipeline Syst. Eng. Pract. 2013.4.

J. Pipeline Syst. Eng. Pract.

0.02
Kn=2

M/ D3

0.016

Kn=1.5
Kn=1

0.012

No
compaction

0.008
0.004

10

100

1000

10000

Sf

Fig. 10. Bending moments at the crown after side-filling (as functions
of Sf and K n )
Fig. 7. Deformed shape for HDPE pipe after backfilling up to the
crown for different values of K n : (a) K n 1; (b) K n 1.5; (c) K n 2

0.002
Kn=2
Kn-1.5

0.0016

M/ D3

Kn=1

0.0012

No compaction

0.0008

0.0004

0
1

10

100

1000

10000

Sf

Fig. 11. Bending moments at the springline after side-filling (as functions of Sf and K n )

0.3
Kn=2

0.25

Fig. 8. Deformed shape for CS pipe after backfilling up to the crown


for different values of K n : (a) K n 1; (b) K n 1.5; (c) K n 2

Kn=1.5
Kn=1

0.2

N/ D2

Downloaded from ascelibrary.org by L & T Institute of Project Management on 12/25/14. Copyright ASCE. For personal use only; all rights reserved.

No compaction

0.15
0.1
0.05
0
1

10

100

1000

10000

Sf

Fig. 12. Thrust at the crown after side-filling (as functions of Sf


and K n )

Fig. 9. Deformed shape for reinforced concrete pipe after backfilling


up to the crown for different values of K n : (a) K n 1; (b) K n 1.5;
(c) K n 2
ASCE

the bending moment remains small. It is also clear that the increase
in K n magnifies the lateral stress and increases the nonuniform
pressure around pipes, and therefore increases the bending moments for pipes having low Sf values. Again, shallow cover (where
compaction stress produces significant behavior) might reasonably
be defined as less than one diameter.
Figs. 12 and 13 show the normalized thrust forces against Sf for
different kneading factors. The results indicate that the imposed
horizontal stress has almost no effect on the thrust forces at the
springline for pipes having different Sf values. The thrust forces

04013004-6

J. Pipeline Syst. Eng. Pract. 2013.4.

J. Pipeline Syst. Eng. Pract.

0.3
0.25

N/ D2

0.2
0.15

Kn=2
0.1

Kn=1.5
Kn=1

0.05

Downloaded from ascelibrary.org by L & T Institute of Project Management on 12/25/14. Copyright ASCE. For personal use only; all rights reserved.

No compaction
0
1

10

100

1000

10000

Sf

Fig. 13. Thrust at the springline after side-filling (as functions of


Sf and K n )

the resulting calculations should be considered to be preliminary,


and confidence is reduced when used for pipe and soil materials,
geometries, and compaction processes beyond this study). The deformations measured when a small vibratory plate compactor was
employed were much smaller than those for the rammer, and it
appears that there is no need to model soil compaction in this case.
Although evidence was provided regarding the performance of the
technique in flexible pipes, further work is needed to measure
changes in thrust and moment in rigid pipes before the modified
compaction model is employed for those structures.
The technique introduced here may be used to explore various
factors associated with compaction of the side-fill, and may assist
in adjusting the design of high-cost structures, particularly if deflection limits for construction are set as part of the design process by
using approaches like that proposed by McGrath et al. (2002).

References
at the crown for pipes increase as a result of the additional horizontal stresses imposed during compaction simulation. For pipes
having high Sf values, the pipes deform under the imposed pressure and the ratio between the thrust forces at the crown and the
springline does not change significantly. For pipes having low
Sf values, the pipes do not deform under the imposed pressure
and the ratio between the thrust forces at the crown and the springline changed significantly.
There are no test data available for rigid pipes to evaluate
whether estimates of increased moment and thrust at the crown
and invert of the pipe are realistic. Further, K n > 1 may not be reasonable for soil adjacent to these stiff structures (as per the discussion in an earlier section). As a result, the use of K n > 1 to assess
the effect of compaction on stress resultants in rigid pipes
(i.e., structures having low Sf ) is not recommended until physical
evidence has been obtained for these structures.

Conclusions
A technique proposed by Taleb and Moore (1999) to model the
effect of soil compaction on pipe response during buried pipe
installation has been examined. A modification factor, K n , was
introduced to account for soil kneading and repetitive cycles of
compaction during placement of the side-fill. A semiempirical technique was used to assign the initial lateral stress without changing
the initial vertical pressure and vertical soil deformation under selfweight. The deformations of five different pipe products were calculated, whichwere compared to measured values reported by
McGrath et al. (1999) for two specific pipe products and measured
values for a 1.5-m-diameter HDPE pipe backfilled with granular
material that was compacted by using a vibrating plate. It was
found that the deformations of those pipes in which soil was compacted by using a rammer were effectively estimated when a kneading factor of 2 was used; therefore, this value is suggested as a
suitable starting point for use in flexible pipe calculations (although

ASCE

ABAQUS V. 6.5 [Computer software]. Pawtucket, RI Habbit, Karlson and


Sonersen, Inc.
Corrugated Steel Pipe Institute (CSPI). (2002). Handbook of steel drainage
and highway construction products, CSPI, Cambridge, Ontario,
Canada.
Duncan, J. M., and Seed, R. B. (1986). Compaction-induced earth pressure under K o -conditions. J. Geotech. Eng., 112(1), 122.
Hambleton, J. P., and Drescher, A. (2009). Modeling wheel-induced
rutting in soils: Rolling. J. Terramech., 46(2), 3547.
Hoeg, K. (1968). Stresses against underground structural cylinders.
J. Soil Mech. Found. Eng. Div., Proc. ASCE, 94(4), 833858.
Hgel, H. M., Henke, S., and Kinzler, S. (2008). High-performance
ABAQUS simulations in soil mechanics. http://www.tu-harburg.de/
gsc/publikationen/download/AUC2008.pdf (Jan. 16, 2009).
Katona, M. G. (1978). Analysis of long-span culverts by the finite element
method. Transportation Research Record 678, Transportation Research Board, Washington, DC, 5966.
Kelm, M., and Grabe, J. (2004). Numerical simulation of the compaction
of granular soils with vibratory rollers. Cyclic behaviour of soils and
liquefaction phenomena, Triantafyllidis, Greece, 661664.
McGrath, T. J., Selig, E. T., Moore, I. D., Webb, M. C., and Taleb, B.
(2002). Recommended specifications for large span culverts. NCHRP
Rep. 473, Transportation Research Board, Washington, DC.
McGrath, T. J., Selig, E. T., Webb, M. C., and Zoladz, G. V. (1999). Pipe
interaction with the backfill envelope. FHWA-RD-98-191, http://www
.pooledfund.org/Document/Download/41 (Jun. 3, 2010).
Moore, I. D. (2001). Culverts and buried pipelines. Chapter 18, Geotechnical and geoenvironmental handbook, R. K. Rowe, ed., Kluwer Academic Publishers, Dordrecht, Netherlands, 541568.
Petersen, D. L., Nelsen, C. R., Li, G., McGrath, T. J., and Kitane, L. (2010).
Recommended design specifications for live load distribution to
buried structures. NCHRP Rep. 647, Transportation Research Board,
Washington, DC.
Taleb, B., and Moore, I. D. (1999). Metal culvert response to earth loading
performance of two-dimensional analysis. Transportation Research
Record 1656, Transportation Research Board, Washington, DC, 2536.
Xia, K. (2011). Finite element modeling of tire/terrain interaction: Application to predicting soil compaction and tire mobility. J. Terramech.,
48(2), 113123.

04013004-7

J. Pipeline Syst. Eng. Pract. 2013.4.

J. Pipeline Syst. Eng. Pract.

Вам также может понравиться